Volume 126, Issue 9 e2021JA029624
Research Article
Open Access

Measurability of the Nonlinear Response of Electron Distribution Function to Chorus Emissions in the Earth's Radiation Belt

M. Hanzelka,

Corresponding Author

M. Hanzelka

Department of Space Physics, Institute of Atmospheric Physics, Czech Academy of Sciences, Prague, Czech Republic

Faculty of Mathematics and Physics, Charles University, Prague, Czech Republic

Correspondence to:

M. Hanzelka,

mha@ufa.cas.cz

Search for more papers by this author
O. Santolík,

O. Santolík

Department of Space Physics, Institute of Atmospheric Physics, Czech Academy of Sciences, Prague, Czech Republic

Faculty of Mathematics and Physics, Charles University, Prague, Czech Republic

Search for more papers by this author
Y. Omura,

Y. Omura

Research Institute for Sustainable Humanosphere, Kyoto University, Uji, Japan

Search for more papers by this author
I. Kolmašová,

I. Kolmašová

Department of Space Physics, Institute of Atmospheric Physics, Czech Academy of Sciences, Prague, Czech Republic

Faculty of Mathematics and Physics, Charles University, Prague, Czech Republic

Search for more papers by this author
First published: 03 September 2021
Citations: 1

Abstract

We conduct test particle simulations to study the perturbations in a hot electron velocity distribution caused by a rising chorus element propagating parallel to the ambient magnetic field in the Earth's outer radiation belt. The wavefield is constructed from the nonlinear growth theory of chorus emissions of Omura (2021, https://doi.org/10.1186/s40623-021-01380-w), with additional considerations about saturation and propagation of the transverse resonant current being applied to model the subpacket structure. Using Liouville's theorem, we trace electrons back in time to reconstruct the evolution of electron velocity distribution at the magnetic equator. The electromagnetic hole created by nonlinear trapping and transport effects appears as a depression in the velocity distribution, aligned with the resonance velocity curve. We analyze the decrease of particle flux in this depression and estimate the energy resolution, pitch angle resolution, time resolution and geometric factor of particle analyzers needed to observe the perturbation. We conclude that particle detectors on current or recently operating spacecraft are always lacking in at least one of these parameters, which explains the missing direct observations of sharp phase space density depressions during chorus-electron nonlinear resonant interaction. However, with a dedicated experiment and appropriate measurement strategy, such observations are within the possibilities of the current technology. Similarity of the simulated density perturbation and a step function mathematical model is used to draw an analogy between the backward wave oscillator regime of chorus generation and the nonlinear growth theory.

Key Points

  • We analyze perturbations in a hot electron distribution caused by nonlinear interactions with a model chorus element with fine structure

  • A stripe structure of phase space density depletions and elevations are observed, associated with individual subpackets

  • Resolution of spacecraft instruments required to observe the leading most prominent stripe is estimated

Plain Language Summary

The plasma environment in the Earth's magnetosphere supports natural growth of various electromagnetic waves, including the whistler-mode chorus emissions, which consist of nonlinear chirping tones. These emissions can reach large amplitudes and play a major role in energization of radiation belt electrons. Nonlinear theories of chorus generation imply microscopic perturbations in resonant electron populations. A long-standing problem is that these predictions were never directly confirmed by experimental observations. Here, we analyze perturbations of electron distribution functions numerically, taking into account spacecraft measurements of short subpackets within each chirping element. We reveal distinct perturbations, which are just below the measurability limits of existing spacecraft instruments. We thus explain the current absence of direct measurements of nonlinear effects of chorus on the electron distribution functions. We also suggest measurement strategies for future spacecraft instruments that can increase the number of detected interaction events.

1 Introduction

The chorus electromagnetic emissions are a class of whistler-mode waves that occur naturally in planetary magnetospheres (Burtis & Helliwell, 1976; Hospodarsky et al., 2008; Tsurutani & Smith, 1974). They occur abundantly in the Earth's radiation belts, mostly in the midnight-to-noon sector (W. Li et al., 2009), and are recognized as an important driver of electron acceleration in this region (Baker et al., 2018; Thorne et al., 2013). They are responsible for the production of the relativistic “killer” electrons (Horne, 2007), as well as for electron precipitation into the upper atmosphere (Bortnik & Thorne, 2007; Ozaki et al., 2019), which creates weak, pulsating auroras (Ma et al., 2020; Miyoshi et al., 2015). The energetic electrons interact with counter-streaming whistler-mode waves through Doppler-shifted cyclotron resonance (Karpman & Shklyar, 1977; Tsurutani & Lakhina, 1997). The energy of the resonant particles ranges typically from about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0001 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0002 for the lower frequency band (0.1–0.5 urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0003, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0004 denotes the local electron gyrofrequency) (W. Li et al., 2010) and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0005 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0006 for the upper frequency band (0.5–0.8 urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0007) (Inan et al., 1992; Meredith et al., 2009). Between those two bands, a power gap is often observed in the chorus spectra (Burtis & Helliwell, 1969; Gao et al., 2019; Santolík et al., 2003), whose origin is possibly related to the formation of two anisotropic hot electron populations that generate each frequency band (Artemyev & Mourenas, 2020; J. Li et al., 2019; Ratcliffe & Watt, 2017).

The lower band part of the emissions often shows highly structured spectra consisting of coherent, high-amplitude discrete elements with rising or falling frequencies (Burtis & Helliwell, 1969; Santolík et al., 20042009). Based on observational evidence and theoretical investigations, the chorus wave elements are known to be formed around the magnetic equator and to propagate bidirectionally to higher latitudes (Demekhov et al., 2020; Lauben et al., 2002; Santolík et al., 2004), approximately following the magnetic field lines (Hanzelka & Santolík, 2019; Santolík, Macúšová, et al., 2014). Observations of the Cluster spacecraft mission revealed a fine structure of chorus emissions with short wave subpackets within each element (Santolík et al., 2003). This fine structure has also been later observed by the Van Allen Probes (Crabtree et al., 2017; Santolík, Kletzing, et al., 2014).

Several theories have been proposed to explain the formation of chorus emissions. In the initial stage of the growth process, unstable anisotropic electron populations transfer energy to whistler-mode waves through the cyclotron resonance (Gołkowski et al., 2019; Kennel & Petschek, 1966; Omura et al., 2008). This mechanism is well described by the linear theory. In the second stage, whistler-mode waves with the highest amplitude start entrapping increasing numbers of resonant electrons, disturbing thus the gyrotropy (uniform distribution in the phase of the gyrating perpendicular velocity vector) of the electron distribution, which causes the formation of resonant currents (Nunn, 1974; Omura et al., 1991; Trakhtengerts et al., 2003). This nongyrotropic distribution carries a depression or elevation in phase space density, which is sometimes called an electromagnetic hole or hill (Omura et al., 2013). The resonant current modifies the dispersion properties of the wave and results in an enhanced, nonlinear amplitude growth, and in a drifting wave frequency. Simplified calculations of the resonant current have been carried out to obtain an estimate of the frequency drift and amplitude growth (Omura et al., 2008; Summers et al., 2012). Other approaches replace the electromagnetic hole with a step function in the parallel velocity distribution function, which then serves as the energy source for the nonlinear growth (Demekhov, 2011; Trakhtengerts, 1999). However, no theoretical approach so far has explained the formation of the subpackets within chorus elements. Observation-based models (Hanzelka et al., 2020) and kinetic particle simulations (Hikishima et al., 2010; Ke et al., 2017; Tao et al., 2017) can reproduce the fine structure, but the fundamental physics behind the subpackets is still unclear.

A mechanism of packet formation through wave beating has been proposed by Santolík (2008). This hypothesis is further supported by the analysis of Crabtree et al. (2017) who identified the simultaneous presence of multiple distinct waves within some chorus elements, especially toward the higher frequency end of the element. New statistical analyses of frequency variation and lengths of subpackets (Zhang, Mourenas, et al., 2020) further support the idea that the formation of fine structure can be at least partially explained by simple superposition of waves, especially in the case of short subpackets, with the longer ones being better explained by trapping-induced amplitude modulation (Tao et al., 2020). Nunn et al. (2021) conducted Vlasov hybrid simulations with two simultaneously occurring elements at similar frequencies (cf., Katoh & Omura, 2016; W. Li et al., 2011) and showed that the combined waveform features a rapid amplitude modulation, consistent with observations of the fine structure of chorus waves. It has been demonstrated with reduced kinetic theory (Mourenas et al., 2018) and by test particle simulations (Hiraga & Omura, 2020) that the presence of fine structure modifies the effectivity of electron acceleration, as the amplitude modulation and random phase jumps between subpackets disturb the trapped particle populations (Zhang, Agapitov, et al., 2020). Detailed understanding of the structure of chorus wave packets is therefore important for accurate predictions of energization and scattering of electrons in the Earth's radiation belts.

Particle analyzers on spacecraft orbiting in the Earth's magnetosphere have provided many new results in the recent years which improved our understanding of nonlinear wave-particle interactions. Fennell et al. (2014) have analyzed Van Allen Probes (RBSP) field and particle data and identified bursts of resonant electrons correlated with upper band chorus activity. Foster et al. (2017) combined theory and RBSP measurements of electron pitch angle distributions to estimate energization of relativistic electrons happening on time scales shorter than a typical chorus element. Perhaps the most direct observation of effects of nonlinear wave-particle interactions was provided by Kurita et al. (2018), who used 8-s measurements of electron fluxes provided by the Arase spacecraft during strong upper band chorus activity and observed transport of resonant electrons in the energy-pitch angle space, which followed the theoretical resonance curves. Nevertheless, spacecraft missions have been unable to detect depressions or elevation in the electron velocity distribution directly correlated with the nonlinear trapping within the duration of a single chorus element. In the related studies of nonlinear resonant interaction between protons and rising tone electromagnetic ion cyclotron (EMIC) waves, Shoji et al. (2017) found significant nongyrotropy in THEMIS proton data associated with occurrence of EMIC risers, which confirms the presence of an electromagnetic proton hole. However, to obtain sufficiently large number of particle counts, they had to average over time intervals containing several elements of the electromagnetic emission.

In this article, we study the perturbations to the hot electron distribution by means of test particle simulations under a fixed chorus wavefield. We include subpacket modulation of the wave amplitude in our model (Section 2) and track the time evolution of the distribution at the magnetic equator. Unlike most of the previous works on this subject, our focus is on obtaining full 2D energy-pitch angle distributions on subpacket timescales. In Sections 3.1 and 3.2, a stripe structure in the velocity distribution is observed and briefly described in agreement with previous kinetic simulations (e.g., Tao et al., 2017). In Section 3.3, we focus on the leading stripe, which is associated with the electromagnetic hole, and analyze the possibility to measure this perturbation with particle analyzers on recently operating spacecraft. Since the stripe in the phase space density has properties of a smoothed step function positioned near the resonance velocity, a successful observation of this feature would provide connection between the backward wave oscillator (BWO) theory of chorus growth, which relies on the existence of such a step function (Demekhov, 2017; Trakhtengerts, 1999), and the nonlinear growth theory of Omura (2021). We conclude that the pitch angle resolution of electron analyzers is insufficient to observe any of these stripes in their entirety. As it turns out, the most crucial deficiency comes from low particle counts on the short timescales of one subpacket, leading to the need of higher geometric factors of particle analyzers, or a change in measurement strategies.

2 Methods

2.1 Chorus Wave Field Model

To get meaningful results from test particle simulations, we need to construct a wavefield model that resembles chorus elements as observed by spacecraft. In Figure 1, we show an example of a rising lower band chorus element, which we used as a reference for our model. The spectrogram of the total magnetic field power spectral density in Figure 1a was obtained from the multicomponent continuous burst-mode measurements of the EMFISIS Waves instrument on the Van Allen Probes B spacecraft. In this mode, waveforms of all six electromagnetic field components were captured at a sampling rate of 35,000 samples per second. The waveforms were analyzed using 512-point Fast Fourier Transform with the Hann window, 50% overlap, and averaging over three neighboring spectra. Calibrated magnetic field waveforms were transformed into the coordinates linked to the background magnetic field (Figure 1b). They were subsequently band-pass filtered between 0.1 and 0.49 of the local electron cyclotron frequency. The instantaneous frequency (Figure 1c) was then obtained as the time derivative of the instantaneous phase of an analytic signal reconstructed using the Hilbert transform. Frequency data from the time intervals where the amplitude dropped below urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0008 were cut off (Santolík, Kletzing, et al., 2014). The plasma frequency urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0009 was obtained from measurements of the upper hybrid frequency (Kurth et al., 2015).

Details are in the caption following the image

A chorus emission as detected by the Van Allen Probes B spacecraft. (a) A power spectrogram obtained from a 6-s burst-mode snapshot of three-component magnetic field waveforms, showing an isolated discrete rising element in the lower frequency band of chorus. Half electron gyrofrequency is represented by a solid white line. (b) 200-ms time interval showing the waveform of the oscillatory magnetic field components parallel (red) and perpendicular (blue) to the local magnetic field line. These waveforms represent measurements from the instrument's full frequency range of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0010 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0011. The subpacket structure of a single chorus element is apparent here, with local maxima shown by black dots. The maximum wave amplitude on this interval is 0.6% of the ambient magnetic field. (c) Instantaneous wave frequency obtained through the Hilbert transform of the magnetic field waveforms filtered to the frequency range depicted by the gray rectangle in panel (a). The overall almost linear trend in the frequency growth is disturbed by irregularities on the subpacket time scale.

The example element is about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0012 long and rises from frequency of about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0013 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0014. After the initial long and weak subpacket, two well separated, high-amplitude packets appear, reaching about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0015 of the background field magnitude. With increasing wave frequency, the subpacket structure becomes less regular. The frequency follows a linear trend with irregularities appearing not only in between, but also inside subpackets, suggesting that this behavior cannot be caused purely by Hilbert transform applied on low-amplitude sections. The wave remains quasiparallel during the measurement, with wave normal angle at peak amplitudes starting at about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0016 and steadily decreasing.

To capture the observed behavior of the depicted chorus element, we use the parallel whistler-mode chorus model of Hanzelka et al. (2020), which is based on the nonlinear growth theory of chorus emissions (Omura, 2021; Omura et al., 2009). This model solves the so-called chorus equations to compute the evolution of plane wave frequency and amplitude in the source of a subpacket, and after the subpacket reaches saturation amplitude and decays, a new subpacket starts at a point in time and space determined by the propagation of the resonant current to upstream regions (Helliwell, 1967; Trakhtengerts et al., 2003). To obtain the electromagnetic field in the downstream, we used transport equations for wave magnetic field urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0017 and angular frequency urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0018 to propagate the wave in one spatial dimension urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0019 along the magnetic field line. These two equations are (Nunn, 1974; Omura et al., 2008)
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0020(1)
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0021(2)

The evolution of the frequency is a pure advection with group velocity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0022, while the amplitude equation is modified by a nonlinear term on the right-hand side which includes the resonant current component urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0023 parallel to the wave electric field. These equations are solved by the upwind method with a time step urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0024 and a spatial step urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0025. On this grid, the Courant-Friedrichs-Lewy condition is satisfied over the whole domain while the numerical diffusion is kept reasonably slow (LeVeque, 2007).

The initial boundary values for each subpacket are given by the coupled set of modified chorus equations (Omura et al., 2009; Hanzelka et al., 2020). For detailed description of these equations, the notation and the complete set of input parameters, see Appendix A, where the results of Hanzelka et al. (2020) are briefly summarized.

To obtain the phase urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0026 of the wave magnetic field vector, frequencies urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0027 and wave numbers urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0028 are integrated for each subpacket (we first integrate over time in the source, then over space away from the source). During particle simulations, these integrated phase angles are bilinearly interpolated at the position of each particle to obtain the angle urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0029 between urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0030 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0031. The phase difference between adjacent subpackets is not specified, as it changes with the movement of the source region.

The final model is presented in Figure 2 as plots of magnetic field amplitudes and wave frequencies. In the following particle simulations, we assume waves propagating in both directions from the magnetic equator, as shown in the amplitude plot in Figure 2b. Figures 2c and 2d show the frequency growth within the subpackets, ranging from urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0032 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0033, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0034 is the equatorial angular gyrofrequency. The simulation was stopped when the starting frequency of the next subpacket would become higher than urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0035; this upper frequency limit corresponds to observations of lower band chorus with a spectral gap, but it is arbitrary in our simulation procedure as there is no natural cutoff for strictly parallel propagating whistler-mode waves. The simulated chorus element features a realistic subpacket structure with an irregular growth of the wave frequency (compare with Figure 1c) and with an upstream shift of the source region—these properties have been observed in numerical simulations (Hikishima et al., 2009) and also in spacecraft measurements (Demekhov et al., 2020; Foster et al., 2017; Santolík, Kletzing, et al., 2014).

Details are in the caption following the image

The chorus wavefield used in our test particle simulations. (a) Time evolution of the wave magnetic field amplitude urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0036 of a chorus element propagating along the magnetic field line in the positive sense of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0037. Plotted for latitudes urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0038 (red line) and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0039 (blue line). (b) Evolution of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0040 in time and space with dotted lines showing the spatial cuts at urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0041 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0042. The total wavefield was obtained as a superposition of the left- and right-propagating waves. Panels (c and d) show the wave frequency urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0043 and copy the format of panels (a and b), with only the right-propagating element being shown. The time duration of the chorus element at the equator is about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0044.

2.2 Backward-in-Time Test Particle Simulation

To obtain a high-resolution velocity distribution at a given point in space and time, we use the backward-in-time simulation method (Nunn & Omura, 2015). In this approach, the equation of motion for electrons in an electromagnetic field is solved from a fixed final point (after the interaction with chorus) to an arbitrary initial point at which the wavefield disappears. The phase space densities corresponding to the velocity of electrons at the end of the simulation are mapped to the fixed final point with the use of Liouville's theorem (Goldstein et al., 2001): the velocity distribution function is constant along the trajectories of a Hamiltonian system. Therefore, if we know the initial, unperturbed distribution at all points along the magnetic field line and we choose the velocities of electrons so that we cover uniformly the velocity space at the final point, we can obtain the phase space density (PSD) of the electron distribution after the interaction with a chorus wave (see Figure 3 for a schematic explanation of the ideas behind the method). This method is obviously not self-consistent, as it requires a precalculated wavefield which does not respond to the changes in the electron velocity distribution. However, unlike in self-consistent kinetic simulations, we can choose to analyze an arbitrary section of the velocity space, thus increasing the resolution. And, maybe even more importantly, the initial distribution can be chosen a posteriori, meaning that we can obtain the perturbations for all possible distributions with just one simulation.

Details are in the caption following the image

A schematic illustration of the particle simulation method. (1) We start the particles from the final point urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0045, where we want to construct the velocity distribution function for electrons. Velocities are chosen so that each particle represents a bin in the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0046 velocity space (in the simulation, the full 3D space urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0047 is used, but the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0048 coordinate is omitted here for the sake of clarity). (2) Particles propagate back in time prior to their interaction with the wavefield to a point urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0049 where the initial PSD is known. (3) Knowing the initial particle velocity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0050, we can read of the value of the distribution function at this point of the velocity space. (4) Employing Liouville's theorem, the value of the distribution function urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0051 is identified with urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0052.

The test particles are advanced backward-in-time by a type of the Boris algorithm (Higuera & Cary, 2017) which does not use the common particle phase approximation (Zenitani & Umeda, 2018). The algorithm preserves phase space volume, which is an important property that allows us to safely apply Liouville's theorem in the numerical simulation. The effect of the mirror force on electrons is included through additional perpendicular components of the background magnetic field (Katoh & Omura, 2006). We also tested that a finer grid in the wave model has only negligible influence on the structure and magnitude of observed perturbations in the electron velocity distribution (peak value of relative PSD change in the simulation with one subpacket changed by less than urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0053). However, some grid-dependent changes can be observed, especially at the edges of electromagnetic holes, as the chaotic trajectory of the resonant electrons strongly depends on their gyrophase at the onset of the resonant interaction.

Since, the phase space density is supposed to be preserved in our physical system, the initial velocity distribution function must evolve adiabatically along field lines. We chose a distribution of the form (Summers et al., 2012)
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0054(3)
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0055(4)
which is a bi-Maxwellian distribution in momenta urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0056, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0057. In velocities urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0058, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0059, the distribution takes form
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0060(5)
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0061(6)

The initial distribution in the particle phases urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0062 is set to be uniform. urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0063 stands here for the number density of hot electrons at the equator, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0064 (or urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0065) and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0066 (or urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0067) are the equatorial thermal velocities (or relativistic momenta) and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0068 is the magnetic dipole field strength along the field line. As noted by Kuzichev et al. (2019), this distribution has a different asymptotic behavior at relativistic velocities than an anisotropic Jüttner distribution. However, we are interested in particles at resonant velocities, which are only weakly relativistic. The loss cone is also not included in the model, because particle behavior at low pitch angles plays no significant role in the nonlinear growth theory.

The input parameters common to all simulation runs are as follows. The ratio of the plasma frequency to the equatorial gyrofrequency is urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0069 (based on measurements shown in Figure 1) and the ratio of the density of hot electrons to the density of cold electrons is urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0070, which is a reasonable value for populations with thermal velocities of a few units to tens of keV (Juhász et al., 2019; Walsh et al., 2020). The parallel and perpendicular thermal relativistic velocities at the equator are urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0071 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0072, resulting in an equatorial temperature anisotropy urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0073. The background magnetic field is assumed to be dipolar with equatorial strength at the Earth's surface of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0074 and the particles are propagating on the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0075-shell with urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0076. The corresponding equatorial electron gyrofrequency evaluates to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0077. These parameters are the same as in the wave model (see Appendix A).

All test particle simulations were done on a uniform grid in (urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0078) space, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0079 is the angle between the wave magnetic field vector and the perpendicular velocity of electrons. The parallel velocities were sampled in the range urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0080 with 256 points, perpendicular velocities in the range urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0081 with 256 points, and phases urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0082 with 64 points (particles with initial velocities larger than urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0083 are excluded from the sample). To confirm that this resolution is sufficient (Voitcu et al., 2012), we limited the range in velocities to focus on the resonance region of the first subpacket, and doubled the number of gyrophase grid points. Apart from some finer features associated with the phase mixing of lower and higher density regions, there was no noticeable difference. In Movies S1–S3, which depict the evolution of phase space hole, the parallel velocities are sampled on a much smaller range and with 1,024 points, the phases are sampled with 512 points and the perpendicular velocities are fixed to a single value. Each frame in these animations requires a new simulation run.

2.3 Particle Count Analysis

Particle count measured by a particle detector with a geometric factor urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0084 over a time interval urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0085 can be expressed as
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0086(7)
where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0087 is the 3D velocity distribution. In our analysis, we work with a velocity distribution averaged over gyrophases, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0088, since it can be safely expected that during the time interval urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0089, the detector will uniformly sample particles across all gyrophases. We define the number flux urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0090 as
urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0091(8)
The width of the azimuthal angle and polar angle of a single pixel of an electrostatic particle detector is included in the geometric factor urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0092, as well as the relative width of a velocity (energy) level urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0093. All detectors listed in Table 1 in Section 3.3 have an approximately constant ratio of the width of the energy bin to the energy at the center of the bin across all levels. The energy geometric factor given in the specification of the detectors is related to the velocity geometric factor as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0094.
Table 1. The Geometric Factor and Resolution in Time, Energy and Polar Angle of Particle Detectors on Selected Spacecraft
FPI HOPE PEACE HEEA MEP-e
MMS RBSP Cluster Arase
Angular resolution (deg) 11.25 18 15 (3.75) 22.5 (3.5)
Energy range (keV) 0.01–30 0.001–50 0.01–26.4 7–87
Energy levels per sweep 32 36 72 16
Time per energy level (ms) 0.23 21 4 15.6
Geometric factor per bin (urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0095) 2.1 0.2 6.0 (1.5) 0.7
  • Note. The listed parameters are based on Pollock et al. (2016) for FPI (Funsten et al., 2013), for HOPE (Johnstone et al., 1997), for PEACE, and on (Kasahara et al., 2018) for medium-energy particle (MEP). The cylindrical MEP instrument has seemingly the best resolution, but it provides only 16 azimuthal channels with large dead zones, thus providing data from urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0096 of each urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0097 channel. For the rest of the instruments, the angular resolution refers to the polar angle bins.

The number of particles required to measure a significant decrease in phase space density can be defined in a variety of ways. Here we use the one-sigma (68%) confidence interval for a Poisson distribution approximated as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0098, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0099 is the number of particles measured on the perturbed distribution in a given energy-angle bin, and we require that the number of particles urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0100 measured in the corresponding bin of the initial (unperturbed) distribution is at least by urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0101 larger. Since relative change in number flux or PSD is directly proportional to the difference in number of measured particles, we get urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0102, where the subscript “0” denotes the initial state. The required number of particles measured in a single bin of the perturbed distribution can be then defined as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0103. If we were to measure an increase in phase space density, we would relate the difference to the higher value of PSD, which would be the perturbed value, so we can generalize the requirement to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0104.

We must also be mindful of the absolute number of measured particles. If we calculate the so-called exact confidence interval (Meeker et al., 2017), it will have different values for different urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0105, for example: urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0106 for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0107, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0108 for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0109, and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0110 for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0111. We can see that our approximated upper bound urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0112 is noticeably underestimated for low urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0113, so we should always require urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0114 in our assessment of measurability to make the approximations valid.

If we decided to work with a two-sigma (95%) confidence interval, the required number of particles would be about four times larger, with the upper bound of the exact confidence interval showing underestimation with nearly the same relative magnitude as in the one-sigma case.

3 Results

3.1 Response of the Electron Distribution to a Single Chorus Subpacket

We start by analyzing the evolution of the equatorial hot electron distribution during interaction with the first subpacket of our wave model from Section 2.1. To numerically simulate the wave-particle interactions, we use backward-in-time test particle simulations as described in Section 2.2. The final point of particle propagation (initial in the backward sense) is urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0115, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0116, lying on the boundary between the first and the second subpacket. The time step is chosen to be a fixed value of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0117. The particles remain near the equator where the magnetic field is almost constant, meaning that each electron gyroperiod is sampled with about 120 points.

During the resonant interaction, an electromagnetic hole forms in the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0118 space (see Movie S1 where the evolution of the electromagnetic hole in the first two subpackets is captured). Particles trapped in the hole are transported to lower parallel velocities, while resonant particles that flow around the hole increase their parallel velocity. For a highly anisotropic distribution, this translates to an increase of the pitch angle urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0119 and the kinetic energy urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0120 for trapped particles, and a decrease of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0121 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0122 for untrapped particles (see Figures 4a and 4b for a sketch of the particle scattering and transport, and see Movies S2 and S3 for changes in the kinetic energy and pitch angle during the propagation of electrons through the first two subpackets). After integrating over particle phases, we obtain a 2D velocity distribution where the hole is not apparent anymore, and the perturbed part of the distribution appears as a stripe along the relativistic resonance velocity curve (see Figure 4c). The resonance velocity is defined as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0123, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0124 is the Lorentz factor and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0125 is the wave number. As expected from the inspection of the electromagnetic hole, the stripe consists of an increased phase space density part which is located close to resonance velocity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0126 corresponding to the initial frequency of the subpacket, and a decreased density part around resonance velocity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0127 corresponding to the wave frequency at the end of the subpacket (see Figure 4d). After integration over perpendicular velocities, we observe a smooth step-like feature in the reduced 1D distribution located around the resonance velocity (see Figure 4e).

Details are in the caption following the image

(a and b) A schematic explanation of the motion of resonant particles. (a) The electromagnetic phase space hole in urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0128 space. The green arrow represents the direction of motion of the trapping region (in gray) to lower urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0129, the red arrow shows the motion of untrapped resonant particles. The streamlines in the phase space are only illustrative and do not represent the full dynamics. (b) Motion of resonant particles illustrated in the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0130 space. Blue and purple lines show the contours of phase space density of an isotropic and a highly anisotropic distribution, respectively (blue lines thus also represent isolines of kinetic energy). Dotted black lines indicate the resonance velocity curves for the lowest and highest frequency within the subpacket. (c–e) Perturbation of the electron velocity distribution after the interaction with one chorus subpacket, simulation starting at point urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0131, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0132. (c) 2D velocity distribution normalized to the maximum phase space density urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0133 at urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0134, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0135. (d) Difference between the perturbed and the initial velocity distribution in urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0136 space, normalized to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0137. (e) The velocity distribution integrated over perpendicular velocities, comparison of the perturbed distribution (blue line) with the initial distribution (red line).

3.2 Response of the Electron Distribution to a Full Chorus Element

In the analysis of the resonant interaction of electrons with a chorus element, we proceed in the same fashion as in the previous case of interaction with a single subpacket. Certain particles may now reach regions (or, in the forward time flow, come from regions) with significantly stronger magnetic field, so the time step must be appropriately smaller. A fixed value of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0138 was chosen, which translates to about 80 steps per electron gyroperiod at the boundary of the simulation domain where the magnetic field is about three times stronger than at the equator. Furthermore, some particles may reflect at their adiabatic mirror point and interact with the chorus element that propagates toward lower urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0139 values.

We set the final point to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0140, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0141, which is right after the end of the last subpacket at the equator. The stripes created by the interaction with the subpackets are now distorted by adiabatic motion of particles that resonated with these subpackets farther from the equator (see Figure 5a; also see Movie S4 for subpacket-by-subpacket evolution of the perturbed distribution at the equator). In the 1D reduced distribution, the stripe-like perturbation related to single subpackets remains visible in those cases where the adiabatic motion of resonant electrons straightened the stripes and aligned them thus with the lines of constant parallel velocity (see Figure 5b). Subtraction of the initial distribution reveals an increase in phase space density at higher parallel velocities and a more prominent density decrease at lower velocities, especially around the resonance velocity corresponding to the last subpacket (see Figure 5d). Further analysis of the perturbations in the 2D distribution reveals that the stripes are overlapping, which makes them hard to distinguish, especially at resonance velocities corresponding to higher frequencies (see Figure 5c; also see the second half of Movie S1 to observe the mixing of phase space density between the first and second subpacket).

Details are in the caption following the image

Perturbation of electron velocity distribution after interaction with urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0142, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0143 (a) 2D reduced velocity distribution showing the stripe-like perturbation due to the subpacket structure of the model chorus element. (b) 1D integrated distribution reveals a decrease of density at lower parallel velocities, especially around the resonance velocity of the last subpacket urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0144. The inset plot shows the detail of this decrease. (c) Normalized difference between perturbed and initial 2D distributions shows the overlapping of stripes and mixing of phase space regions, leading to a distortion of the stripe structure. The transverse region of diminished perturbations at high perpendicular velocities is due to particles reflected at their mirror points and is not essential for our study (see Figures S1 and S2 for a visual explanation). (d) Normalized difference between the perturbed and the initial 1D distribution. The stripe structure is still visible in the integrated distribution, with the low-velocity decrease of about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0145 being the most prominent feature.

3.3 Analysis of the Required Resolution of Particle Measurements

The first step in this measurability analysis is to determine the time interval and the range of energies and pitch angles in which we observe large phase space density perturbations and, at the same time, large particle fluxes. If we assume the number urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0146 of particles measured in each energy-angle bin to be Poisson distributed, then the statistical uncertainty of the measurement can be represented by the relative standard deviation urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0147. So, in order to make the measurement of a PSD perturbation significant at the one-sigma level, we need at least urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0148 particles in each bin, where urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0149 is the number flux (see Section 2 for a more detailed explanation of confidence intervals and the definition of number flux). This means that unlike in Figure 4d, we need to look at the relative perturbations of the phase space density and associated fluxes. The relative change in PSD after interaction with the first subpacket is shown in Figure 6a. The values of the number flux urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0150 shown in Figure 6b, which are proportional to the particle counts, reach maximum at urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0151, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0152. The quantity we use to determine the most suitable phase space region for measurements is urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0153. A plot of this quantity is presented in Figure 6c for the case of the first subpacket. By looking at the state of the perturbed distribution after interaction with each subpacket, we determined that the highest values of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0154 are found between subpackets 4 and 7. The case of the fifth subpacket is shown in Figure 6d. It needs to be stressed here that in our nonself-consistent simulation, we cannot determine how fast will the stripe on the left (associated with the first subpacket) decay, and so we should focus only on the newly formed generated right-hand stripe where the nonlinear interaction is taking place.

Details are in the caption following the image

(a) Relative change in 2D velocity distribution after interaction with one chorus subpacket. Same data as in Figures 4c–4e. (b) Number flux of electrons in the highly anisotropic initial velocity distribution, given in normalized units. (c) The ratio of the number flux and the required particle count for the case of interaction with the first subpacket. Contours of constant energy are plotted with a step of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0155, with each 50 keV level highlighted with a thicker line. (d) Same type of plot as panel (c), but the distribution was captured after interaction with the fifth subpacket. The black arrows point at the phase space region where the subpacket generation is currently taking place.

From Figure 6d we conclude that the optimal energy range for measurement of the phase space density decrease associated with growth of chorus subpackets starts at around urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0156 and ends at about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0157. In this energy range, the pitch angle width of the stripe is about 6°. A detector that always resolves the stripe should have a pitch angle resolution of half this width, which is 3°. Because the stripe does not lie on a constant pitch angle, the width of energy bins is also limited, but does not pose a major constraint.

In the second step of the measurability analysis, we compare the size and magnitude of the simulated perturbations with the resolution of an actual spacecraft instrument. Table 1 lists electrostatic particle detectors on several currently or recently operating spacecraft with science objectives related to wave-particle interactions in the magnetosphere. We can see that only the medium-energy particle-electrons (MEP-e) particle analyzer on the Arase spacecraft (Kasahara et al., 2018) has a suitable energy range, so we used this detector as a reference in further analysis. The MEP-e is a disklike electron spectrometer which filters incoming electrons by energy with an electrostatic analyzer and then uses avalanche photodiodes for particle detection. The time resolution of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0158 is close to the time duration of one chorus subpacket, meaning that we can use data from only one energy level to avoid smearing of the density stripe we wish to observe. With 16 logarithmically spaced energy levels from urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0159 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0160, only the highest energy level is suitable to observe the structure depicted in Figure 6d (marked by a black arrow). The FPI detector on the MMS spacecraft and PEACE on the Cluster spacecraft provide better time resolution, but it compromises the particle counts they can obtain. For the purpose of comparison with our simulation, we assumed that the detector is oriented in such a way that it covers full urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0161 in the pitch angle. Thus, the angular resolution in Table 1 is taken to be equivalent to pitch angle bins, although better resolution could be achieved with a different orientation and a more limited range of pitch angles. Changes of effectivity of the detector as a function of energy are not significant and are not considered here.

Since the measuring capabilities of electron analyzers on spacecraft are restricted to recording the particle count as a function of their kinetic energy and pitch angles, we will from now on inspect the perturbations of the electron distribution using these variables. Furthermore, all aforementioned simulation results represent the distribution at a single instant in time, while spacecraft particle analyzers typically require tens to hundreds of milliseconds to sweep through the relevant range of electron energies. To mimic the particle instrument behavior, we performed 16 backward particle simulations with the final time point being at the end of the fifth subpacket, and in each simulation, the final time point moved back by urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0162, adding up to a total time interval of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0163. The chosen energy bin from urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0164 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0165 was covered with 16 logarithmically spaced grid points and we used 256 linearly spaced points for the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0166 pitch angle quadrant (64 points per single MEP-e angular sector) and 128 linearly spaced points for particle velocity phases.

Unfortunately, due to the large dead zones of the azimuthal channels of the MEP-e analyzer, it is immediately clear that we cannot resolve the perturbation in flux, because we have only one urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0167 angular bin per a urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0168 channel (see Table 1). Therefore, we had to base our simulation on a hypothetical detector which preserves the parameters of MEP-e, but has more azimuthal channels and negligible dead zones. We proceeded to refine the pitch angle bins to find out a compromise between high particle counts and good angular resolution. The urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0169 ratio fluctuated as we were refining the angular grid, depending on whether the energy-angle bin fell into the center of the stripe or on the edge. However, with bin size urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0170 or smaller, the ratio was growing steadily, signifying that by further resizing the bins, we are only losing particle counts and not improving the resolution of the stripe anymore. For bins of size between urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0171 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0172, the ratio fluctuated between 1 and 2, with expected number of particles falling down to 3 or 4 per bin in the region of the PSD decrease. To achieve reasonably large particle counts, and also to prevent errors in estimation of confidence intervals, we further had to increase the geometric factor 10 times with respect to the original pitch angle resolution (i.e., the effective aperture of the detector has been increased 10 times), making it comparable to the PEACE detector.

In Figures 7a–7c, we can see an example with angular bin size of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0173. A decrease in phase space density is observed between urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0174, followed by fluctuations at larger values of pitch angle. Thus, with the increased geometric factor and angular resolution, the phase space density stripe can be resolved, and the particle counts shown in Figure 7b are above 50. As shown in Figure 7c, the required particle counts for one-sigma significance are met at the lowest point of the dip. We can reach even urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0175, corresponding to a two-sigma significance. The points where the nonlinear wave growth is taking place are marked by green diamonds in the plot.

Details are in the caption following the image

Pitch angle distributions of phase space density and particle counts. (a) Normalized difference between perturbed and initial pitch angle distribution at a single energy level. Steps in pitch angle of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0176 are chosen as the best compromise between good angular resolution and high particle counts, the geometric factor is 10 times that of MEP-e. Energy level and accumulation period are based of MEP-e. (b) Number of particles expected to be measured on the perturbed distribution. The numbers for the initial distribution are indicated with the dotted line. (c) Particle counts urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0177 required to make the normalized standard deviation of Poisson distribution equal to the relative change in the PSD. Ratio urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0178 is plotted with a red line. The green diamonds mark pitch angle bins in which the PSD decrease associated with the last subpacket is observed. Time resolution and geometric factor used in the calculation match the MEP-e parameters shown in Table 1. (d) Normalized difference between perturbed and initial PSD in kinetic energies and pitch angles. Ratio between energy levels is the same as in the MEP-e detector, but the range is extended, and the improved angular resolution and geometric factor from panels (a-c) are used. Region with the PSD decrease has a black borderline. (e) Expected particle counts. (f) Ratio of required particle counts to expected particle counts.

To get a more complete picture of the stripe as it could be seen by particle detectors, we further assumed four detectors operating simultaneously at four logarithmically spaced energy levels from urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0179 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0180. We preserved the accumulation time step and energy level spacing as given in the MEP-e specifications, and we used the improved pitch angle resolution of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0181 and the increased geometric factor of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0182 as in Figures 7a-7c above. As expected from the above analysis, the PSD decrease is well resolved in both energies and pitch angles and the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0183 ratio reaches values much smaller than one at each energy level.

We conclude that with the particle detectors mounted on currently or recently operating spacecraft represented by Arase, Van Allen Probes, Cluster and MMS, a direct measurement of perturbations in electron velocity distribution due to interaction with a chorus subpackets is not possible, mainly due to insufficient angular resolution and low upper energy limits. With a dedicated experiment, however, an in situ observation of these perturbations should be within the current technical possibilities. Such an experiment would consist of particle analyzers measuring electrons with energies up to at least urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0184 with a polar angle resolution better than about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0185, and with a geometric factor of at least urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0186 (i.e., 10-times better than MEP-e). Unlike with FPI on MMS, obtaining a fast 3D distribution would not be of essence: during high chorus wave activity, the analyzer would operate in a fixed high energy mode with accumulation times comparable to the theoretical time duration of a single subpacket. We believe that with this type of experiment, it should be possible to obtain direct confirmation of the processes described by the nonlinear growth theory of chorus rising tone emissions.

4 Summary and Discussion

We have studied the nonlinear interactions between a rising-frequency chorus element with a subpacket structure and a hot electron population. Each subpacket traps resonant electrons from a low phase space density part of the anisotropic velocity distribution and transports them to higher energies and pitch angles, while untrapped resonant particles experience scattering to lower energies and pitch angles. In result, we observe stripes of increased and decreased phase space density forming along the elliptical resonance curve, with increased PSD at the resonance corresponding to the initial frequency, and decreased PSD at the resonance corresponding to the terminal frequency of the subpacket. With each subsequent subpacket, some of the electrons from the phase space density depletion become trapped again. Due to this consecutive trapping, a very prominent depletion is formed at the position of the leading subpacket as the system evolves. Since we are observing the evolution from the equator, the perturbations which formed at higher latitudes will arrive to the observation point with adiabatic correction to pitch angle. Therefore, the stripes associated with previous subpackets do not follow resonance curves anymore, with their high pitch angle parts being shifted toward larger parallel velocities.

We further considered the possibility to measure the prominent PSD depletion by spacecraft particle instruments. Such measurements require energy and pitch angle range which covers the part of the depletion where the maximum change to particle flux is expected. Moreover, accumulation times are limited to a window in the order of tens of milliseconds due to the fast propagation of the perturbation through the velocity space. It turns out that none of the available instruments meets the criteria for a statistically significant observation. This negative result can be seen as an explanation for the lack of direct observational evidence of these perturbations. Nevertheless, we showed that with the energy resolution of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0187 near the energy level of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0188 (similar to the MEP-e instrument on Arase), a pitch angle resolution of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0189 (similar to MEP-e, but without dead zones), and with a geometric factor of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0190 (similar to that of the PEACE HEEA instrument on Cluster), it would be possible to achieve a conclusive measurement results on the phase space density depletion.

In our test particle simulation, we have used a fixed model of the wavefield of the chorus element. Such nonself-consistent approach can yield a realistic evolution of the electron velocity distribution function only if the model well represents the actual chorus wave fields present in the radiation belts. As it is technically impossible to directly measure a large time-space section of the wavefield, we can make a comparison only with kinetic simulations. Looking at some recent particle-in-cell (PIC) simulations (Hikishima et al., 2010; Tao et al., 2017), we can confirm that the phase space density decrease related to the last subpacket is well apparent in their results, while the rest of the stripe structure is only marginally visible. Our hypothesis here is that the stripes might decay while generating a broadband whistler-mode spectrum through a wave growth mechanism driven by increased pitch angle anisotropies along the stripes. Because this part of the spectrum is not present in our wave model, the additional stripes cannot decay in the test particle simulation. Based on the proposed hypothesis and the available PIC simulation results, we believe that observable effects of chorus subpackets on electron distributions primarily occur on the leading stripe with a large phase space density decrease. We note here that due to the sequential nature of our wave model and the drifting source assumption, each subpacket represents a rising-frequency plane wave with its phase independent of the previous packet, and so the limiting effect of random phase jumps on nonlinear electron acceleration (Zhang, Agapitov, et al., 2020) is naturally included. It should be also mentioned that the multi-wave structure (Crabtree et al., 2017) of chorus elements leads to higher variability in amplitude and packet length than assumed in our simple model, which could result in even more pronounced overlaps between resonance widths of each subpacket and add further smearing of the stripe structure. Unfortunately, the lack of observational evidence for the stripe structure cannot be used as evidence for multi-wave models, since this structure shows even smaller perturbation amplitudes than the PSD decrease analyzed in Section 3.3.

The presented analysis of measurability included some simplifications, mainly by assuming a special orientation of the electron analyzers with respect to the ambient magnetic field. A different inclination of the spectrometer could limit the total range of pitch angles by removing intervals around urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0191 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0192. This would be acceptable since the interesting part of the distribution lies approximately within urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0193 from the perpendicular direction. Also, we referenced mainly electrostatic analyzers, but given the required energy range and the advantage of a fixed energy operation mode, semiconductor detectors similar to RAPID (Wilken et al., 2001) on Cluster or FEEPS (Blake et al., 2016) on MMS might be more suitable for this task. However, the software on both RAPID and FEEPS was designed to provide data in time steps which are orders of magnitude above the subpacket time scale, so they were not considered in this paper.

We further note that the wavefield we used was based on a chorus event observed in region with a very low plasma-to-gyrofrequency ratio urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0194, and a very large perpendicular thermal velocity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0195 was assumed as well. These choices increase the resonance velocity and shift fluxes toward higher energies. A different wavefield with urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0196 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0197 was tested, which resulted in the optimal urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0198 values starting at about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0199, which is still outside the range of most of the available electrostatic analyzers.

Another simplification we used is the limitation to a single model of hot plasma distribution. Based on previous statistical results (W. Li et al., 2010), the phase space density of 10 – urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0200 electrons measured between urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0201 and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0202 near the equator ranges from about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0203 to urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0204 during geomagnetically disturbed conditions. In our model, the average density in this energy range is about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0205 at urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0206, and so assuming even larger densities to achieve higher particle counts would make our estimates unrealistic. Further concerning the hot plasma distribution, we assume that it is smooth in its initial state. Given that there is a continuous presence of whistler waves in the radiation belts, occurrence of some fluctuations in the phase space density should be expected. The coherent nature of chorus will always lead to organization of the electrons and to creation of the electromagnetic hole, giving us confidence that the leading stripe structure will still be prominent, but the stripes outside of the active growth region might be further distorted by the ambient fluctuations.

The obtained phase space density decrease at the leading subpacket is similar to the step-like perturbation assumed in the BWO (Backward Wave Oscillator) chorus growth theory (Demekhov, 2017). However, the step in the 1D distribution (see the inset plot in Figure 5b) is not very sharp due to the relativistic effects included in the resonance velocity. Relativistic versions of the BWO theory formulas would be needed to make a comparison with our results. As this phase space density perturbation should be observable with a dedicated particle experiment, the same experiment should be also able to confirm experimentally the existence of the (relativistic) step function from the BWO theory. Positive outcome from the experiment would show a clear relation between the step function paradigm and the electromagnetic hole approach. The recent work of Zonca et al. (2021) presents a general theoretical framework for chorus excitation which hints at a fundamental connection between gyrotron backwave oscillators and other interpretations of chorus chirping (Omura & Nunn, 2011; Vomvoridis et al., 1982), supporting thus further experimental and theoretical research in this area.

In conclusion, our simulations show that the electron distribution function carries an imprint of the fine structure of nonlinear wave packets within chirping chorus elements. Spacecraft observation of the full stripe structure appearing in our simulated distribution would complement the research of subpacket formation in chorus elements by providing information about resonant currents, but unfortunately, such measurements are not feasible. On the other hand, the major perturbation associated with the front of the element provides connection between the BWO theory and the nonlinear growth theory of Omura (2021), and the possible detection of such perturbation is just beyond the possibilities of the currently available instrumentation. Better geometric factors of energetic electron analyzers are needed, along with operational modes focused on gathering large particle counts in relevant energy and pitch angle ranges. These improvements are thus needed to further support the validity of current theories of the chorus growth mechanism.

Notation

  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0207
  • equatorial temperature anisotropy.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0208
  • amplitude of wave magnetic field.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0209
  • speed of light in vacuum.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0210
  • elementary charge.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0211
  • position of the source of the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0212-th subpacket along the field line.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0213
  • resonant current component parallel with the wave electric field.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0214
  • resonant current component parallel with the wave magnetic field.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0215
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0216 computed for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0217.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0218
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0219 computed for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0220.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0221
  • wave number.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0222
  • electron mass.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0223
  • hot electron number density.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0224
  • depth of the electron phase space hole.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0225
  • factors entering the calculation of the inhomogeneity urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0226.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0227
  • inhomogeneity ratio (Omura et al., 2008).
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0228
  • negative value of the inhomogeneity ratio in the source.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0229
  • parallel relativistic particle momentum urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0230.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0231
  • perpendicular relativistic particle momentum urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0232.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0233
  • parallel thermal relativistic velocity, with subscript “eq” denoting the equatorial value.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0234
  • perpendicular thermal relativistic velocity, with subscript “eq” denoting the equatorial value.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0235
  • group velocity of whistler-mode wave.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0236
  • cyclotron resonance velocity.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0237
  • typical perpendicular velocity of particles, appears in the nonlinear growth theory in the function urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0238 which replaces the perpendicular factor of the electron velocity distribution function.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0239
  • Lorentz factor.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0240
  • Lorentz factor for urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0241.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0242
  • nonlinear growth rate.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0243
  • difference between wave magnetic field phase and perpendicular particle velocity phase.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0244
  • vacuum permeability.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0245
  • wave angular frequency.
  • urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0246
  • normalized wave amplitude urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0247.
  • Acknowledgments

    This work has received funding from the European Union's Horizon 2020 research and innovation programme under grant agreement No. 870452 (PAGER). M. Hanzelka, I. Kolmasova, and O. Santolik further acknowledge support from the Czech Academy of Sciences through the Mobility Plus grant JSPS-19-05, and from the MEYS grant LTAUSA17070. M. Hanzelka and O. Santolik acknowledge support from Charles University through the GA UK project No. 64120. Y. Omura acknowledges support from the JSPS KAKENHI Grant 17H06140.

      Appendix A: Chorus Wave Model Equations

      To calculate the chorus wavefield which is at the background of all particle simulations conducted in this study, we use the model of Hanzelka et al. (2020) with corrections based on Omura (2021). In the model it is assumed that the source of each subpacket is a point in space, specifically urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0248 for the first subpacket. The amplitude at urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0249 is double the threshold amplitude
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0250(A1)
      Here and thereafter, the common parabolic approximation urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0251 is not used since we simulate the wavefield propagation in an exact dipole magnetic field. The time evolution of the wave frequency and amplitude in the source is described by the coupled equations
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0252(A2)
      and
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0253(A3)
      where
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0254(A4)
      is the nonlinear growth rate. When the optimum amplitude
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0255(A5)
      is reached, the sign of the amplitude growth is switched in order to simulate saturation and decrease in amplitude. A new subpacket is assumed to be triggered by a residual resonant current at a point urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0256 from which the wave would propagate (according to cold plasma dispersion) to the exact point where the amplitude of the previous subpacket drops below urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0257. The propagation of wave amplitude and frequency in space and time is described by the advection Equations 2 and 1. The resonant current component parallel with the wave electric field is given by
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0258(A6)
      where
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0259(A7)
      and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0260, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0261 are given by the shape of the electron phase space hole (Omura et al., 2008). The hot electron distribution function enters the calculation through the quantity
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0262(A8)
      We obtained urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0263 through methods described by Summers et al. (2012) using a parabolic approximation of the Earth's dipole magnetic field model and the distribution function
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0264(A9)
      with the hot electron density
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0265(A10)
      and the perpendicular thermal velocity
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0266(A11)

      The average momentum urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0267 at distance urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0268 is related to the thermal momentum as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0269. The distribution in Equation A9 is the same one as in our particle simulations.

      In the work of Hanzelka et al. (2020), each new subpacket was let to evolve independently of the previous subpacket. But the electron hole structure which produces the resonant current has a certain width in the velocity space, and so if two waves that experience the nonlinear growth are overlapping, they need to comply to the frequency separability criterion (Omura et al., 2015, Equations 36–39). In our model, each new subpacket is triggered at such a point in time and space that during its propagation, it does not collide with the source of the previous packet in the time-space diagram. However, the model includes a certain overlap of frequency ranges of each two adjacent subpacket. Therefore, during the downstream propagation, adjacent packets will inevitably start merging due to the difference in group velocities. We assume that in the overlapping region, the production of the resonant current in the new subpacket is supressed, thus limiting the convective growth prescribed by Equation 2. In practice we multiply the resonant current calculated at point urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0270 of the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0271-th subpacket, overlapping with the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0272-th subpacket, by a factor
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0273(A12)
      where
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0274(A13)
      urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0275 is the estimated frequency bandwidth corresponding to the trapping potential (Omura et al., 2015) and urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0276 is the frequency of the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0277-th subpacket. This means that the resonant current in the urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0278-th subpacket is progressively more suppressed when the frequency separation urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0279 drops below the limiting bandwidth of urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0280. In Figure 2a we can see that with this suppression of current, the wave amplitudes urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0281 reach about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0282 at the equator and increase by less than a factor of 2 after reaching magnetic latitude urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0283. The first subpacket is an outlier as the simulation proceeds sequentially, subpacket-by-subpacket, and thus we cannot apply the factor urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0284 to the first wave. However, the convective growth results in a maximum amplitude of about urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0285 (Figure 2b), which is comparable to some of the most intense whistler waves observed in the inner magnetosphere (Cattell et al., 2008). Also, the wave amplitudes inside chorus elements tend to be the highest in the first few subpackets – see for example, Figure 4b in Santolík, Kletzing, et al. (2014), or Figure 1b in this text.

      Finally, the evolution equations are solved by the upwind method with time step urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0286 and spatial step urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0287. The equatorial strength of the dipole field at the surface of the Earth is chosen as urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0288, L-value of the field line is urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0289. Further parameters are chosen as follows: urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0290, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0291, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0292, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0293, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0294, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0295, urn:x-wiley:21699380:media:jgra56730:jgra56730-math-0296.

      Data Availability Statement

      The Van Allen Probe data are publicly available from the NASA's Space Physics Data Facility, repository https://spdf.gsfc.nasa.gov/pub/data/rbsp/. Data obtained from numerical simulations are available for download at https://doi.org/10.6084/m9.figshare.14315936.v1.